41
HAL Id: hal-01428947 https://hal.archives-ouvertes.fr/hal-01428947 Submitted on 11 Jan 2017 HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers. L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés. Metamodel-based sensitivity analysis: polynomial chaos expansions and Gaussian processes Loic Le Gratiet, Stefano Marelli, Bruno Sudret To cite this version: Loic Le Gratiet, Stefano Marelli, Bruno Sudret. Metamodel-based sensitivity analysis: polynomial chaos expansions and Gaussian processes. Handbook of Uncertainty Quantification - Part III: Sensi- tivity analysis, 2016. <hal-01428947>

Metamodel-based sensitivity analysis: polynomial … · Metamodel-based sensitivity analysis: polynomial ... Metamodel-based sensitivity analysis: polynomial chaos expansions

  • Upload
    dohanh

  • View
    245

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Metamodel-based sensitivity analysis: polynomial … · Metamodel-based sensitivity analysis: polynomial ... Metamodel-based sensitivity analysis: polynomial chaos expansions

HAL Id: hal-01428947https://hal.archives-ouvertes.fr/hal-01428947

Submitted on 11 Jan 2017

HAL is a multi-disciplinary open accessarchive for the deposit and dissemination of sci-entific research documents, whether they are pub-lished or not. The documents may come fromteaching and research institutions in France orabroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, estdestinée au dépôt et à la diffusion de documentsscientifiques de niveau recherche, publiés ou non,émanant des établissements d’enseignement et derecherche français ou étrangers, des laboratoirespublics ou privés.

Metamodel-based sensitivity analysis: polynomial chaosexpansions and Gaussian processesLoic Le Gratiet, Stefano Marelli, Bruno Sudret

To cite this version:Loic Le Gratiet, Stefano Marelli, Bruno Sudret. Metamodel-based sensitivity analysis: polynomialchaos expansions and Gaussian processes. Handbook of Uncertainty Quantification - Part III: Sensi-tivity analysis, 2016. <hal-01428947>

Page 2: Metamodel-based sensitivity analysis: polynomial … · Metamodel-based sensitivity analysis: polynomial ... Metamodel-based sensitivity analysis: polynomial chaos expansions

METAMODEL-BASED SENSITIVITY ANALYSIS:POLYNOMIAL CHAOS EXPANSIONS AND GAUSSIAN

PROCESSES

L. Le Gratiet, S. Marelli, B. Sudret

CHAIR OF RISK, SAFETY AND UNCERTAINTY QUANTIFICATION

STEFANO-FRANSCINI-PLATZ 5CH-8093 ZURICH

Risk, Safety &Uncertainty Quantification

Page 3: Metamodel-based sensitivity analysis: polynomial … · Metamodel-based sensitivity analysis: polynomial ... Metamodel-based sensitivity analysis: polynomial chaos expansions

Data Sheet

Journal: Handbook of Uncertainty Quantification - Part III: Sensitivity analysis

Report Ref.: RSUQ-2016-007

Arxiv Ref.: http://arxiv.org/abs/1606.04273 - [stat.CO]

DOI: -

Date submitted: January 2016

Date accepted: March 2016

Page 4: Metamodel-based sensitivity analysis: polynomial … · Metamodel-based sensitivity analysis: polynomial ... Metamodel-based sensitivity analysis: polynomial chaos expansions

Metamodel-based sensitivity analysis: Polynomial chaos

expansions and Gaussian processes

Loıc Le Gratiet1, Stefano Marelli2, and Bruno Sudret2

1EDF R&D, 6 quai Watier, 78401 Chatou, France

2 ETH Zurich, Chair of Risk, Safety & Uncertainty Quantification, Stefano-Franscini-Platz 5, CH-8093 Zurich,

Switzerland

Abstract

Global sensitivity analysis is now established as a powerful approach for determining the key

random input parameters that drive the uncertainty of model output predictions. Yet the classical

computation of the so-called Sobol’ indices is based on Monte Carlo simulation, which is not af-

fordable when computationally expensive models are used, as it is the case in most applications in

engineering and applied sciences. In this respect metamodels such as polynomial chaos expansions

(PCE) and Gaussian processes (GP) have received tremendous attention in the last few years, as they

allow one to replace the original, taxing model by a surrogate which is built from an experimental

design of limited size. Then the surrogate can be used to compute the sensitivity indices in negligible

time. In this chapter an introduction to each technique is given, with an emphasis on their strengths

and limitations in the context of global sensitivity analysis. In particular, Sobol’ (resp. total Sobol’)

indices can be computed analytically from the PCE coefficients. In contrast, confidence intervals on

sensitivity indices can be derived straightforwardly from the properties of GPs. The performance of

the two techniques is finally compared on three well-known analytical benchmarks (Ishigami, G-Sobol

and Morris functions) as well as on a realistic engineering application (deflection of a truss structure).

Keywords: Polynomial Chaos Expansions, Gaussian Processes, Kriging, Error estimation, Sobol’

indices

1 Introduction

In modern engineering sciences computational models are used to simulate and predict the behavior of

complex systems. The governing equations of the system are usually discretized so as to be solved by

dedicated algorithms. In the end a computational model (a.k.a. simulator) is built up, which can be

considered as a mapping from the space of input parameters to the space of quantities of interest that

are computed by the model. However, in many situations the values of the parameters describing the

properties of the system, its environment and the various initial and boundary conditions are not perfectly

1

Page 5: Metamodel-based sensitivity analysis: polynomial … · Metamodel-based sensitivity analysis: polynomial ... Metamodel-based sensitivity analysis: polynomial chaos expansions

well-known. To account for such uncertainty, they are typically described by possible variation ranges or

probability distribution functions.

In this context global sensitivity analysis aims at determining which input parameters of the model

influence the most the predictions, i.e. how the variability of the model response is affected by the

uncertainty of the various input parameters. A popular technique is based on the decomposition of the

response variance as a sum of contributions that can be associated to each single input parameter or to

combinations thereof, leading to the computation of the so-called Sobol’ indices.

As presented earlier in this book (see Variance-based sensitivity analysis: Theory and estimation algo-

rithms), the use of Monte Carlo simulation to compute Sobol’ indices requires a large number of samples

(typically, thousands to hundreds of thousands), which may be an impossible requirement when the un-

derlying computational model is expensive-to-evaluate. To bypass this difficulty, surrogate models may

be built. Generally speaking, a surrogate model (a.k.a. metamodel or emulator) is an approximation of

the original computational model:

x ∈ DX ⊂ Rd 7→ y = G(x) (1)

which is constructed based on a limited number of runs of the true model, the so-called experimental

design:

X =x(1), . . . ,x(n)

. (2)

Once a type of surrogate model is selected, the parameters have to be fitted based on the information

contained in the experimental design X and associated runs of the original computational model Y =yi = G(x(i)), i = 1, . . . , n

. Then the accuracy of the surrogate shall be estimated by some kind of

validation technique. For a general introduction to surrogate modelling the reader is referred to [55] and

to the recent review by Iooss and Lemaıtre [29].

In this chapter we discuss two classes of surrogate models that are commonly used for sensitivity analysis,

namely polynomial chaos expansions (PCE) and Gaussian processes (GP). The use of polynomial chaos

expansions in the context of sensitivity analysis has been originally presented in Sudret [56, 58] using a

non intrusive least-square method. Other non-intrusive strategies for the calculation of PCE coefficients

include spectral projection through sparse grids (e.g. Crestaux et al. [19]; Buzzard and Xiu [16]; Buzzard

[15]) and sparse polynomial expansions (e.g. Blatman and Sudret [12]). In the last five years numerous

application examples have been developed using PCE for sensitivity analysis, e.g. Fajraoui et al. [25];

Younes et al. [65]; Brown et al. [14]; Sandoval et al. [45]. Recent extensions to problems with dependent

input parameters can be found in Sudret and Caniou [60]; Munoz Zuniga et al. [40].

In parallel, Gaussian process modeling has been introduced in the context of sensitivity analysis by Welch

et al. [63]; Oakley and O’Hagan [42]; Marrel et al. [38, 37]. Recent developments in which metamodeling

errors are taken into account in the analysis have been proposed by Le Gratiet et al. [34]; Chastaing and

Le Gratiet [17].

The chapter first recalls the basics of the two approaches and details how they can be used to compute

sensitivity indices. The two approaches are then compared on different benchmark examples as well as

on an application in structural mechanics.

2

Page 6: Metamodel-based sensitivity analysis: polynomial … · Metamodel-based sensitivity analysis: polynomial ... Metamodel-based sensitivity analysis: polynomial chaos expansions

2 Polynomial chaos expansions

2.1 Mathematical setup

Let us consider a computational model G : x ∈ DX ⊂ Rd 7→ y = G(x) ∈ R. Suppose that the

uncertainty in the input parameters is modeled by a random vector X with prescribed joint probability

density function (PDF) fX(x). The resulting (random) quantity of interest Y = G(X) is obtained by

propagating the uncertainty inX through G. Assuming that Y has a finite variance (which is a physically

meaningful assumption when dealing with physical systems), it belongs to the so-called Hilbert space of

second order random variables, which allows for the following spectral representation to hold [53]:

Y =

∞∑

j=0

yj Zj . (3)

The random variable Y is therefore cast as an infinite series, in which Zj∞j=0 is a numerable set of

random variables (which form a basis of the Hilbert space), and yj∞j=0 are coefficients. The latter may

be interpreted as the coordinates of Y in this basis. In the sequel we focus on polynomial chaos expansions,

in which the basis terms Zj∞j=0 are multivariate orthonormal polynomials in the input vector X, i.e.

Zj = Ψj(X).

2.2 Polynomial chaos basis

In the sequel we assume that the input variables are statistically independent, so that the joint PDF is

the product of the d marginal distributions: fX(x) =∏di=1 fXi

(xi), where the fXi(xi) are the marginal

distributions of each variable Xi, i = 1, . . . , d defined on DXi. For each single variable Xi and any

two functions φ1, φ2 : x ∈ DXi 7→ R, we define the functional inner product by the following integral

(provided it exists):

〈φ1, φ2〉i def= E [φ1(Xi)φ2(Xi)] =

DXi

φ1(x)φ2(x) fXi(x) dx. (4)

Using the above notation, classical algebra allows one to build a family of orthogonal polynomials P (i)k , k ∈

N satisfying

⟨P

(i)j , P

(i)k

⟩i

def= E

[P

(i)j (Xi)P

(i)k (Xi)

]=

DXi

P(i)j (x) P

(i)k (x) fXi

(x) dx = a(i)j δjk, (5)

see e.g. Abramowitz and Stegun [1]. In the above equation subscript k denotes the degree of the

polynomial P(i)k , δjk is the Kronecker symbol equal to 1 when j = k and 0 otherwise and a

(i)j corresponds

to the squared norm of P(i)j :

a(i)j

def=‖ P (i)

j ‖2i =⟨P

(i)j , P

(i)j

⟩i. (6)

In general orthogonal bases may be obtained by applying the Gram-Schmidt orthogonalization proce-

dure, e.g. to the canonical family of monomials

1, x, x2, . . .

. For standard distributions, the associated

families of orthogonal polynomials are well-known [64]. For instance, ifXi ∼ U(−1, 1) has a uniform distri-

bution over [−1, 1], the resulting family is that of the so-called Legendre polynomials. When Xi ∼ N (0, 1)

3

Page 7: Metamodel-based sensitivity analysis: polynomial … · Metamodel-based sensitivity analysis: polynomial ... Metamodel-based sensitivity analysis: polynomial chaos expansions

has a standard normal distribution with zero mean value and unit standard deviation, the resulting family

is that of the Hermite polynomials. The families associated to standard distributions are summarized in

Table 1 (taken from Sudret [57]).

Table 1: Classical families of orthogonal polynomials (taken from Sudret [57])

Type of variable Distribution Orthogonal polynomials Hilbertian basis ψk(x)

Uniform

U(−1, 1)1[−1,1](x)/2 Legendre Pk(x) Pk(x)/

√1

2k+1

Gaussian

N (0, 1)1√2πe−x

2/2 Hermite Hek(x) Hek(x)/√k!

Gamma

Γ(a, λ = 1)xa e−x 1R+(x) Laguerre Lak(x) Lak(x)/

√Γ(k+a+1)

k!

Beta

B(a, b)1[−1,1](x) (1−x)a(1+x)b

B(a)B(b) Jacobi Ja,bk (x) Ja,bk (x)/Ja,b,k

J2a,b,k = 2a+b+1

2k+a+b+1Γ(k+a+1)Γ(k+b+1)Γ(k+a+b+1)Γ(k+1)

Note that the obtained family is usually not orthonormal. By enforcing normalization, an orthonormal

familyψ

(i)j

∞j=0

is obtained from Eqs.(5),(6) as follows (see Table 1):

ψ(i)j = P

(i)j /

√a

(i)j i = 1, . . . , d, j ∈ N. (7)

From the sets of univariate orthonormal polynomials one can now build multivariate orthonormal polyno-

mials with a tensor product construction. For this purpose let us define the multi-indices α ∈ Nd, which

are ordered lists of integers:

α = (α1, . . . , αd) , αi ∈ N. (8)

One can associate a multivariate polynomial Ψα to any multi-index α by

Ψα(x)def=

d∏

i=1

ψ(i)αi

(xi), (9)

where the univariate polynomialsψ

(i)k , k ∈ N

are defined above, see Eqs.(5),(7). By virtue of Eq.(5)

and the above tensor product construction, the multivariate polynomials in the input vector X are also

orthonormal, i.e.

E [Ψα(X) Ψβ(X)]def=

DX

Ψα(x)Ψβ(x) fX(x) dx = δαβ ∀α,β ∈ Nd, (10)

where δαβ is the Kronecker symbol which is equal to 1 if α = β and zero otherwise. With this notation,

it can be proven that the set of all multivariate polynomials in the input random vector X forms a basis

of the Hilbert space in which Y = G(X) is to be represented [53]:

Y =∑

α∈Nd

yαΨα(X). (11)

4

Page 8: Metamodel-based sensitivity analysis: polynomial … · Metamodel-based sensitivity analysis: polynomial ... Metamodel-based sensitivity analysis: polynomial chaos expansions

2.3 Non standard variables and truncation scheme

In practical sensitivity analysis problems the input variables may not necessarily have standardized dis-

tributions as the ones described in Table 1. Thus reduced variables U with standardized distributions are

introduced first through an isoprobabilistic transform:

X = T (U). (12)

For instance, when dealing with independent uniform distributions with support DXi = [ai, bi], i =

1, . . . , d, the isoprobabilistic transform reads:

Xi =ai + bi

2+bi − ai

2Ui Ui ∼ U([−1, 1]). (13)

In the case of Gaussian independent variables Xi ∼ N (µi , σi) , i = 1, . . . , d, the one-to-one mapping

reads:

Xi = µi + σi Ui, Ui ∼ N (0, 1) (14)

In the general case when the input variables are non Gaussian (e.g. Gumbel distributions, see application

in Section 4.4), the one-to-one mapping may be obtained as follows:

Xi = F−1Xi

(Φ(Ui)) Ui ∼ N (0, 1) (15)

where FXi(resp. Φ) is the cumulative distribution function (CDF) of variable Xi (resp. the standard

normal CDF).

This isoprobabilistic transform approach also allows one to address problems with dependent variables.

For instance, if the input vector X is defined by a set of marginal distributions and a Gaussian copula,

it can be transformed into a set of independent standard normal variables using the Nataf transform

[20; 35].

The representation of the random response in Eq.(11) is exact when the infinite series is considered.

However, in practice, only a finite number of terms may be computed. For this purpose a truncation

scheme has to be selected. Since the polynomial chaos basis consists of multivariate polynomials, it is

natural to consider as a truncated series all the polynomials up to a given maximum degree. Let us define

the total degree of a multivariate polynomial Ψα by:

|α| def=

d∑

i=1

αi. (16)

The standard truncation scheme consists in selecting all polynomials such that |α| is smaller than or equal

to a given p. This leads to a set of polynomials denoted by Ad,p =α ∈ Nd : |α| ≤ p

of cardinality:

card Ad,p =

(d+ p

p

)=

(d+ p)!

d! p!. (17)

The maximal polynomial degree p may typically be equal to 3 − 5 in practical applications. Note that

the cardinality of Ad,p increases exponentially with d and p. Thus the number of terms in the series,

i.e. the number of coefficients to be computed, increases dramatically when d is large, say d > 10. This

complexity is referred to as the curse of dimensionality. This issue may be solved using specific algorithms

to compute sparse PCE, see e.g. Blatman and Sudret [13]; Doostan and Owhadi [21].

5

Page 9: Metamodel-based sensitivity analysis: polynomial … · Metamodel-based sensitivity analysis: polynomial ... Metamodel-based sensitivity analysis: polynomial chaos expansions

2.4 Computation of the coefficients and error estimation

The use of polynomial chaos expansions has emerged in the late eighties in uncertainty quantification

problems under the form of stochastic finite element methods [26]. In this setup the constitutive equa-

tions of the physical problem are discretized both in the physical space (using standard finite element

techniques) and in the random space using polynomial chaos expansion. This results in coupled systems

of equations which require ad-hoc solvers, thus the term “intrusive approach”.

Non intrusive techniques such as projection or stochastic collocation have emerged in the last decade as

a means to compute the coefficients of PC expansions from repeated evaluations of the existing model G

considered as a black-box function. In this section we focus on a particular non intrusive approach based

on least-square analysis.

Following Berveiller et al. [6, 7], the computation of the PCE coefficients may be cast as a least-square

minimization problem (originally termed “regression” problem) as follows: once a truncation scheme

A ⊂ Nd is chosen (for instance, A = Ad,p), the infinite series is recast as the sum of the truncated series

and a residual:

Y = G(X) =∑

α∈AyαΨα(X) + ε, (18)

in which ε corresponds to all those PC polynomials whose index α is not in the truncation set A. The

least-square minimization approach consists in finding the set of coefficients y = yα, α ∈ A which

minimizes the mean square error

E[ε2] def

= E

(G(X)−

α∈AyαΨα(X)

)2 . (19)

The set of coefficients y is computed at once by solving:

y = arg miny∈RcardA

E

(G(X)−

α∈AyαΨα(X)

)2 . (20)

In practice the discretized version of the problem is obtained by replacing the expectation operator in

Eq.(20) by the empirical mean over a sample set:

y = arg miny∈RcardA

1

N

N∑

i=1

(G(x(i))−

α∈AyαΨα(x(i))

)2

. (21)

In this expression, X =x(i), i = 1, . . . , n

is a sample set of points called experimental design (ED)

that is typically obtained by Monte Carlo simulation of the input random vector X. To solve the least-

square minimization problem in Eq.(21) the computational model G is first run for each point in the ED,

and the results are stored in a vector Y =y(1) = G(x(1)), . . . , y(n) = G(x(n))

T. Then the so-called

information matrix is calculated from the evaluation of the basis polynomials onto each point in the ED:

A =Aij

def= Ψj(x

(i)) , i = 1, . . . , n, j = 1, . . . , card A. (22)

The solution of the least-square minimization problem finally reads:

y =(ATA

)−1

AT Y. (23)

6

Page 10: Metamodel-based sensitivity analysis: polynomial … · Metamodel-based sensitivity analysis: polynomial ... Metamodel-based sensitivity analysis: polynomial chaos expansions

The points used in the experimental design may be obtained from crude Monte Carlo simulation. However

other types of designs are of common use, especially Latin Hypercube sampling (LHS), see McKay et al.

[39], or quasi-random sequences such as the Sobol’ or Halton sequence [41]. The size of the experimental

design is of crucial importance: it must be larger than the number of unknowns cardA for the problem

to be well-posed. In practice we use the thumb rule n ≈ 2 - 3 card A [9].

The simple least-square approach summarized above does not allow one to cope with the curse of dimen-

sionality. Indeed the standard truncation scheme requires approximately 3 ·(d+pp

)runs of the original

model G(x), which is in the order of 104 when e.g. d ≥ 15, p ≥ 5. However, in practice most of the

problems lead eventually to sparse expansions, i.e. PCE in which most of the coefficients are zero or

negligible. In order to find directly the significant polynomials and associated coefficients, sparse PCE

have been introduced recently by Blatman and Sudret [10, 11]; Bieri and Schwab [8]. The recent develop-

ments make use of specific selection algorithms which, by solving a penalized least-square problem, lead

by construction to sparse expansions. Of interest in this chapter is the use of the least-angle regression

algorithm (LAR, Efron et al. [24]), which was introduced in the field of uncertainty quantification by

Blatman and Sudret [13]. Details can be found in Sudret [59]. Note that other techniques based on

compressive sensing have also been developed recently, see e.g. Doostan and Owhadi [21]; Sargsyan et al.

[47]; Jakeman et al. [30].

2.5 Error estimation

The truncation of the polynomial chaos expansion introduces an approximation error which may be

computed a posteriori. Based on the data contained in the experimental design, the empirical error may

be computed from Eq.(21) once least-square minimization problem has been solved:

εemp =1

N

N∑

i=1

(G(x(i))−

α∈AyαΨα(x(i))

)2

. (24)

However, this estimator usually underestimates severely the mean square error in Eq.(19). In particular,

if the size N of the experimental design is close to the number of unknown coefficients card A, the

empirical error tends to zero whereas the true mean square error does not.

A more robust error estimator can be derived based on the cross-validation technique. The experimental

design is split into a training set and a validation set: the coefficients of the expansion are computed

using the training set (Eq.(21)) whereas the error is estimated using the validation set. The leave-one-out

cross-validation is a particular case in which all points but one are used to compute the coefficients.

Setting aside x(i) ∈ X , a PCE denoted by GPC\i(X) is built up using the experimental design X\x(i) def=

x(1), . . . ,x(i−1), x(i+1), . . . ,x(n)

. Then the error is computed at point x(i):

∆idef= G(x(i))−GPC\i(x(i)). (25)

The LOO error is defined by:

εLOO =1

n

n∑

i=1

∆2i =

1

n

n∑

i=1

(G(x(i))−GPC\i(x(i))

)2

. (26)

7

Page 11: Metamodel-based sensitivity analysis: polynomial … · Metamodel-based sensitivity analysis: polynomial ... Metamodel-based sensitivity analysis: polynomial chaos expansions

After some algebra this reduces to:

εLOO =1

n

n∑

i=1

(G(x(i))−GPC(x(i))

1− hi

)2

, (27)

where hi is the i-th diagonal term of matrix A(ATA)−1AT (matrix A is defined in Eq.(22)) and GPC(·) is

now the PC expansion built up from the full experimental design X .

As a conclusion, when using a least-square minimization technique to compute the coefficients of a PC

expansion, an a posteriori estimator of the mean-square error is readily available. This allows one to

compare PCEs obtained from different truncation schemes and select the best one according to the

leave-one-out error estimate.

2.6 Post-processing for sensitivity analysis

2.6.1 Statistical moments

The truncated PC expansion Y = GPC(X) =∑α∈A yαΨα(X) contains all the information about the

statistical properties of the random output Y = G(X). Due to the orthogonality of the PC basis, mean

and standard deviation of Y may be computed directly from the coefficients y. Indeed, since Ψ0 ≡ 1, we

get E [Ψα(X)] = 0 ∀α 6= 0. Thus the mean value of Y is the first term of the series:

E[Y]

= E

[∑

α∈AyαΨα(X)

]= y0. (28)

Similarly, due to Eq.(10) the variance of Y may be cast as:

σ2Y

def= Var

[Y]

= E[(Y − y0

)2]

=∑

α∈Aα 6=0

y2α. (29)

In other words the mean and variance of the random response may be obtained by a mere combination

of the PCE coefficients once the latter have been computed.

2.6.2 Sobol’ decomposition and indices

As already discussed in Chapter 4, global sensitivity analysis is based on Sobol’ decomposition of the

computational model G (a.k.a generalized ANOVA decomposition), which reads [50]:

G(x) = G0 +d∑

i=1

Gi(xi) +∑

1≤i<j≤dGij(xi, xj) + · · ·+G12...d(x), (30)

that is, as a sum of a constant G0, univariate functions Gi(xi) , 1 ≤ i ≤ d, bivariate functions

Gij(xi, xj) , 1 ≤ i < j ≤ d, etc. A recursive construction is obtained by the following recurrence rela-

tionship:

G0 = E [G(X)]

Gi(xi) = E [G(X)|Xi = xi]−G0

Gij(xi, xj) = E [G(X)|Xi, Xj = xi, xj ]−Gi(xi)−Gj(xj)−G0.

(31)

8

Page 12: Metamodel-based sensitivity analysis: polynomial … · Metamodel-based sensitivity analysis: polynomial ... Metamodel-based sensitivity analysis: polynomial chaos expansions

Using the set notation for indices

Adef= i1, . . . , is ⊂ 1, . . . , d , (32)

the Sobol’ decomposition in Eq.(30) reads:

G(x) = G0 +∑

A⊂1, ... ,dA6=∅

GA(xA), (33)

where xA is a subvector of x which only contains the components that belong to the index set A. It can

be proven that the summands are orthogonal with each other:

E [GA(xA)GB(xB)] = 0 ∀ A,B ⊂ 1, . . . , d , A 6= B. (34)

Using this orthogonality property, one can decompose the variance of the model output

Vdef= Var [Y ] = Var

A⊂1, ... ,dA6=∅

GA(xA)

=

A⊂1, ... ,dA 6=∅

Var [GA(XA)] (35)

as the sum of so-called partial variances defined by:

VAdef= Var [GA(XA)] = E

[G2A(XA)

]. (36)

The Sobol’ index attached to each subset of variables Adef= i1, . . . , is ⊂ 1, . . . , d is finally defined

by:

SA =VAV

=Var [GA(XA)]

Var [Y ]. (37)

First-order Sobol’ indices quantify the portion of the total variance V that can be apportioned to the

sole input variable Xi:

Si =ViV

=Var [Gi(Xi)]

Var [Y ]. (38)

Second-order indices quantify the joint effect of variables (Xi, Xj) that cannot be explained by each single

variable separately:

Sij =VijV

=Var [Gij(Xi, Xj)]

Var [Y ]. (39)

Finally, total Sobol’ indices Stoti quantify the total impact of a given parameter Xi including all of its

interactions with other variables. They may be computed by the sum of the Sobol’ indices of any order

that contain Xi:

Stoti =

A3iSA. (40)

Amongst other methods, Monte Carlo estimators of the various indices are available in the literature

and thoroughly discussed in (see Variance-based sensitivity analysis: Theory and estimation algorithms).

Their computation usually requires 103−4 runs of the model G for each index, which leads to a global

computational cost that is not affordable when G is expensive-to-evaluate.

9

Page 13: Metamodel-based sensitivity analysis: polynomial … · Metamodel-based sensitivity analysis: polynomial ... Metamodel-based sensitivity analysis: polynomial chaos expansions

2.6.3 Sobol’ indices and PC expansions

As can be seen by comparing Eqs.(18) and (33), both polynomial chaos expansions and Sobol’ decom-

position are sums of orthogonal functions. Taking advantage of this property, it is possible to derive

analytic expressions for Sobol’ indices based on a PC expansion, as originally shown in Sudret [56, 58].

For this purpose let us consider the set of multivariate polynomials Ψα which depend only on a subset

of variables A = i1, . . . , is ⊂ 1, . . . , d:

AA = α ∈ A : αk 6= 0 if and only if k ∈ A . (41)

The union of all these sets is by construction equal to A. Thus we can reorder the terms of the truncated

PC expansion so as to exhibit the Sobol’ decomposition:

GPC(x) = y0 +∑

A⊂1, ... ,dA6=∅

GPCA (xA) where GPC

A (xA)def=

α∈AA

yαΨα(x). (42)

Consequently, due to the orthogonality of the PC basis, the partial variance VA reduces to:

VA = Var[GPCA (XA)

]=∑

α∈AA

y2α. (43)

In other words, from a given PC expansion, the Sobol’ indices at any order may be obtained by a mere

combination of the squares of the coefficients. More specifically, the PC-based estimator of the first-order

Sobol’ indices read:

Si =

α∈Ai

y2α

α∈A , α 6=0

y2α

where Ai = α ∈ A : αi > 0 , αj 6=i = 0 . (44)

and the total PC-based Sobol’ indices read:

Stoti =

α∈Atoti

y2α

α∈A , α 6=0

y2α

Atoti = α ∈ A : αi > 0 . (45)

2.7 Summary

Polynomial chaos expansions allow one to cast the random response G(X) as a truncated series expansion.

By selecting an orthonormal basis w.r.t. the input parameter distributions, the corresponding coefficients

can be given a straightforward interpretation: the first coefficient y0 is the mean value of the model output

whereas the variance is the sum of the squares of the remaining coefficients. Similarly, the Sobol’ indices

are obtained by summing up the squares of suitable coefficients. Note that in low dimension (d < 10)

the coefficients can be computed by solving a mere ordinary least-square problem. In higher dimensions

advanced techniques leading to sparse expansions must be used to keep the total computational cost

(measured in terms of the size N of the experimental design) affordable. Yet the post-processing to get

the Sobol’ indices from the PCE coefficients is independent of the technique used.

10

Page 14: Metamodel-based sensitivity analysis: polynomial … · Metamodel-based sensitivity analysis: polynomial ... Metamodel-based sensitivity analysis: polynomial chaos expansions

3 Gaussian process-based sensitivity analysis

3.1 A short introduction to Gaussian processes

Let us consider a probability space (ΩZ ,FZ ,PZ), a measurable space (S,B((S)) and an arbitrary set T .

A stochastic process Z(x), x ∈ T , is Gaussian if and only if for any finite subset C ⊂ T , the collection

of random variables Z(C) has a Gaussian joint distribution. In our framework, T and S represent the

input and the output spaces. Therefore, we have T = Rd and S = R.

A Gaussian process is entirely specified by its mean m(x) = EZ [Z(x)] and covariance function k(x,x′) =

covZ(Z(x), Z(x′)) where EZ and covZ denote the expectation and the covariance with respect to (ΩZ ,FZ ,PZ).

The covariance function k(x,x′) is a positive definite kernel. It is often considered stationary i.e. k(x,x′)

is a function of x−x′. The covariance kernel is the most important term of a Gaussian process regression.

Indeed, it controls the smoothness and the scale of the approximation. A popular choice for k(x,x′) is

the stationary isotropic squared exponential kernel defined as :

k(x,x′) = σ2exp

(− 1

2θ2 ||x− x′||2).

It is parametrized by the parameter θ – also called characteristic length scale or correlation length –

and the variance parameter σ2. We give in Figure 1 examples of realizations of Gaussian processes with

stationary isotropic squared exponential kernels.

We observe that m(x) is the trend around which the realizations vary, σ2 controls the range of their

variation and θ controls their oscillation frequencies. We highligh that Gaussian processes with squared

exponential covariance kernels are infinitely differentiable almost surely. As mentioned in [54], this choice

of kernel can be unrealistic due to its strong regularity.

3.2 Gaussian process regression models

The principle of Gaussian process regression is to consider that the prior knowledge about the computa-

tional model G(x), x ∈ Rd, can be modeled by a Gaussian process Z(x) with a mean denoted by m(x)

and a covariance kernel denoted by k(x,x′). Roughly speaking, we consider that the true response is a

realization of Z(x). Usually, the mean and the covariance are parametrized as follows:

m(x) = fT(x)β, (46)

and

k(x,x′) = σ2r(x,x′;θ), (47)

where fT(x) is a vector of p prescribed functions and β, σ2 and θ have to be estimated. The mean function

m(x) describes the trend and the covariance kernel k(x,x′) describes the regularity and characteristic

length scale of the model.

11

Page 15: Metamodel-based sensitivity analysis: polynomial … · Metamodel-based sensitivity analysis: polynomial ... Metamodel-based sensitivity analysis: polynomial chaos expansions

0.0 0.2 0.4 0.6 0.8 1.0

−3

−2

−1

01

23

σ = 1 ; θ = 0.1 ; m(x) = 0

x

Z(x

)

0.0 0.2 0.4 0.6 0.8 1.0

−4

−2

02

4

σ = 2 ; θ = 0.05 ; m(x) = 0

x

Z(x

)

0.0 0.2 0.4 0.6 0.8 1.0

−3

−2

−1

01

23

σ = 1 ; θ = 0.1 ; m(x) = − 1 + 2x

x

Z(x

)

Figure 1: Examples of Gaussian process realizations with squared exponential kernels and different means.

The shaded areas represent the point-wise 95% confidence intervals.

3.2.1 Predictive distribution

Consider an experimental design X =x(1), . . . ,x(n)

, x(i) ∈ Rd, and the corresponding model responses

Y = G(X ). The predictive distribution of G(x) is given by:

[Z(x)|Z(X ) = Y, σ2,θ] ∼ GP (mn(x), kn(x,x′)) , (48)

where

mn(x) = fT(x)β + rT(x)R−1(Y − Fβ

), (49)

kn(x,x′) = σ2

1−

(fT(x) rT(x)

)0 FT

F R

f(x′)

r(x′)

, (50)

12

Page 16: Metamodel-based sensitivity analysis: polynomial … · Metamodel-based sensitivity analysis: polynomial ... Metamodel-based sensitivity analysis: polynomial chaos expansions

0.0 0.2 0.4 0.6 0.8 1.0

−0.

50.

00.

51.

01.

52.

0

x

f(x)

Figure 2: Examples of predictive distribution. The solid line represents the mean of the predictive

distribution, the non-solid lines represent some of its realizations and the shaded area represents the 95%

confidence intervals based on the variance of the predictive distribution.

In these expressions R = [r(xi,xj ;θ)]i,j=1,...,n, r(x) = [r(x,x(i);θ)]i=1,...,n, F = [fT(x(i))]i=1,...,n and

β =(FTR−1F

)−1FTR−1Y. (51)

The term β denotes the posterior distribution mode of β obtained from the improper non-informative

prior distribution π(β) ∝ 1 [44].

Remark. The predictive distribution is given by the Gaussian process Z(x) conditioned by the known ob-

servations Y. The Gaussian process regression metamodel is given by the conditional expectation mn(x)

and its mean squared error is given by the conditional variance kn(x,x). An illustration of mn(x) and

kn(x,x) is given in Figure 2.

The reader can note that the predictive distribution (48) integrates the posterior distribution of β.

However, the hyper-parameters σ2 and θ are not known in practice and shall be estimated with the

maximum likelihood method [28; 46] or a cross-validation strategy [3]. Then, their estimates are plugged

in the predictive distribution. The restricted maximum likelihood estimate of σ2 is given by:

σ2 =(Y − Fβ)TR−1(Y − Fβ)

n− p . (52)

Unfortunately, such a closed form expression does not exist for θ and it has to be numerically estimated.

Remark. Gaussian process regression can easily be extended to the case of noisy observations. Let us

suppose that Y is tainted by a white Gaussian noise ε :

Yobs = Y + σε(X )ε.

13

Page 17: Metamodel-based sensitivity analysis: polynomial … · Metamodel-based sensitivity analysis: polynomial ... Metamodel-based sensitivity analysis: polynomial chaos expansions

The term σε(X ) represents the standard deviation of the observation noise. The mean and the covariance

of the predictive distribution [Z(x)obs|Z(X ) = Yobs, σ2,θ] is then obtained by replacing in Equations (49),

(50) and (51) the correlation matrix R by σ2R + ∆ε where ∆ε is a diagonal matrix given by :

∆ε =

σε(x(1))

σε(x(2))

. . .

σε(x(n))

.

We emphasize that the closed form expression for the restricted maximum likelihood estimate of σ2 does

not exist anymore. Therefore, this parameter has to be numerically estimated.

3.2.2 Sequential design

To improve the global accuracy of the GP model, it is usual to augment the initial design set X with

new points. An important feature of Gaussian process regression is that it provides an estimate of the

model mean-square error through the term kn(x,x′) (50) which can be used to select these new points.

The most common but not efficient sequential criterion consists in adding the point x(n+1) where the

mean-square error is the largest:

x(n+1) = arg maxx

kn(x,x). (53)

More efficient criteria can be found in Bates et al. [4]; van Beers and Kleijnen [62]; Le Gratiet and

Cannamela [33].

3.2.3 Model selection

To build up a GP model, the user has to make several choices. Indeed, the vector of functions f(x) and

the class of the correlation kernel r(x,x′;θ) need to be set (see Rasmussen and Williams [43] for different

examples of correlation kernels). These choices and the relevance of the model are tested a posteriori with

a validation procedure. If the number n of observations is large, an external validation may be performed

on a test set. Otherwise, a cross-validation procedure may be used. An interesting property of GP mod-

els is that a closed form expression exists for the cross-validation predictive distribution, see for instance

Dubrule [23]. It allows for deriving efficient methods of parameter estimation [3] or sequential design [33].

Some usual stationary covariance kernel are listed below.

The squared exponential covariance function. The form of this kernel is given by:

k(x,x′) = σ2exp

(− 1

2θ2 ||x− x′||2).

This covariance function corresponds to Gaussian processes which are infinitely differentiable in

mean square and almost surely. We illustrate in Figure 3 the 1-dimensional squared exponential

kernel with different correlation lengths and examples of resulting Gaussian process realizations.

14

Page 18: Metamodel-based sensitivity analysis: polynomial … · Metamodel-based sensitivity analysis: polynomial ... Metamodel-based sensitivity analysis: polynomial chaos expansions

0.0 0.2 0.4 0.6 0.8 1.0

0.0

0.2

0.4

0.6

0.8

1.0

h

k(h)

θ = 0.1θ = 0.2θ = 0.3

0.0 0.2 0.4 0.6 0.8 1.0

−2

−1

01

2

xZ(x

)

θ = 0.1θ = 0.2θ = 0.3

Figure 3: The squared exponential kernel in function of h = x − x′ with different correlation lengths θ

and examples of resulting Gaussian process realizations.

The ν-Matern covariance function. This covariance kernel is defined as follow (see [54]):

kν(h) =21−ν

Γ(ν)

(√2||h||θ

)νKν

(√2ν||h||θ

),

where ν is the regularity parameter, Kν is a modified Bessel Function and Γ is the Euler Gamma

function. A Gaussian process with a ν-Matern covariance kernel is ν-Holder continuous in mean

square and ν′-Holder continuous almost surely with ν′ < ν. Three popular choice of ν-Matern

covariance kernels are the ones for ν = 1/2, ν = 3/2 and ν = 5/2 :

kν=1/2(h) = exp

(−||h||θ

),

kν=3/2(h) =

(1 +

√3||h||θ

)exp

(−√

3||h||θ

),

and

kν=5/2(h) =

(1 +

√5||h||θ

+5||h||23θ2

)exp

(−√

5||h||θ

).

We illustrate in Figure 4 the 1-dimensional ν-Matern kernel for different values of ν.

The γ-exponential covariance function. This kernel is defined as follow:

kγ(h) = exp

(−( ||h||θ

)γ).

For γ < 2 the corresponding Gaussian process are not differentiable in mean square whereas for

γ = 2 is is infinitely differentiable (it corresponds to the squared exponential kernel). We illustrate

in Figure 5 the 1-dimensional γ-exponential kernel for different values of γ.

15

Page 19: Metamodel-based sensitivity analysis: polynomial … · Metamodel-based sensitivity analysis: polynomial ... Metamodel-based sensitivity analysis: polynomial chaos expansions

0.0 0.2 0.4 0.6 0.8 1.0

0.0

0.2

0.4

0.6

0.8

1.0

h

k(h)

ν = 1 2ν = 3 2ν = 5 2

0.0 0.2 0.4 0.6 0.8 1.0

−2

−1

01

2

xZ(x

)

ν = 1 2ν = 3 2ν = 5 2

Figure 4: The ν-Matern kernel in function of h = x − x′ with different regularity parameters ν and

examples of resulting Gaussian process realizations.

0.0 0.2 0.4 0.6 0.8 1.0

0.0

0.2

0.4

0.6

0.8

1.0

h

k(h)

γ = 1γ = 3 2γ = 2

0.0 0.2 0.4 0.6 0.8 1.0

−2

−1

01

2

x

Z(x

)

γ = 1γ = 3 2γ = 2

Figure 5: The γ-exponetial kernel in function of h = x − x′ with different regularity parameters γ and

examples of resulting Gaussian process realizations.

3.2.4 Sensitivity analysis

To perform a sensitivity analysis from a GP model, two approaches are possible. The first one consists

in substituting the true model G(x) with the mean of the conditional Gaussian process mn(x) in (49).

However, it may provide biased sensitivity index estimates. Furthermore it does not allow one to quantify

the error on the sensitivity indices due to the metamodel approximation. The second one consists in

16

Page 20: Metamodel-based sensitivity analysis: polynomial … · Metamodel-based sensitivity analysis: polynomial ... Metamodel-based sensitivity analysis: polynomial chaos expansions

substituting G(x) by a Gaussian process Zn(x) having the predictive distribution [Z(x)|Z(X ) = Y, σ2,θ]

shown in (48). This approach makes it possible to quantify the uncertainty due to the metamodel

approximation and allows for building unbiased index estimates.

3.3 Main effects visualization

From now on, the input parameter x ∈ Rd is considered as a random input vector X = (X1, . . . , Xd)

with independent components. Before focusing on variance-based sensitivity indices, the inference about

the main effects is studied in this section. Main effects are a powerful tool to visualize the impact of

each input variable on the model output (see e.g. Oakley and O’Hagan [42]). The main effect of the

group of input variables XA, A ⊂ 1, . . . , d is defined by E [G(X)|XA]. Since the original model G may

be time-consuming to evaluate, it is substituted for by its approximation, i.e. G(X) ≈ E [Zn(X)|XA],

where Zn(x) ∼ [Z(x)|Z(X ) = Y, σ2,θ]. Since E [Zn(X)|XA] is a linear transformation of the Gaussian

process Zn(x), it is also a Gaussian process. The expectations, variances and covariances with respect to

the posterior distribution of [Z(x)|Z(X ) = Y, σ2,θ] are denoted by EZ [.], VarZ (.) and CovZ (., .). Then,

we have:

E [Zn(X)|XA] ∼ GP (E [mn(X)|XA] ,E [E [kn(X,X′)|XA] |X′A]) . (54)

The term E [mn(X)] represents the approximation of E [G(X)|XA] and E [E [kn(X,X′)|XA] |X′A] is the

mean-square error due to the metamodel approximation. Therefore, with this method one can quantify

the error on the main effects due to the metamodel approximation. For more detail about this approach,

the reader is referred to Oakley and O’Hagan [42]; Marrel et al. [37].

3.4 Variance of the main effects

Although the main effect enables one to visualize the impact of a group of variables on the model output,

it does not quantify it. To perform such an analysis, consider the variance of the main effect:

VA = Var (E [Zn(X)|XA]) , (55)

or its normalized version which corresponds to the Sobol’ index:

SA =VAV

=Var (E [Zn(X)|XA])

Var (Zn(X)). (56)

Sobol’ indices are the most popular measures to carry out a sensitivity analysis since their value can

easily be interpreted as the part of the total variance due to a group of variables. However, in contrary to

the partial variance VA, it does not provide information about the order of magnitude of the contribution

to the model output variance of variable group XA.

3.4.1 Analytic formulae

The above indices are studied in Oakley and O’Hagan [42] where the estimation of VA and V is performed

separately. Indeed, computing the Sobol’ index SA requires considering the joint distribution of VA and

17

Page 21: Metamodel-based sensitivity analysis: polynomial … · Metamodel-based sensitivity analysis: polynomial ... Metamodel-based sensitivity analysis: polynomial chaos expansions

V , which makes it impossible to derive analytic formulae. According to Oakley and O’Hagan [42], closed

form expressions in terms of integrals can be obtained for the two quantities EZ [VA] and VarZ (VA). The

quantity EZ [VA] is the sensitivity measure and VarZ (VA) represents the error due to the metamodel

approximation. Nevertheless, VA is not a linear transform of Zn(X) and its full distribution cannot be

established.

3.4.2 Variance estimates with Monte-Carlo integration

To evaluate the Sobol’ index SA, it is possible to use the pick-freeze approaches presented in Chapter 4

and in Sobol [50]; Sobol et al. [52]; Janon et al. [31]. By considering the formula given in Sobol [50], SA

can be approximated by:

SA,N =

1N

∑Ni=1 Zn(X(i))Zn(X

(i)A)−

(1

2N

∑Ni=1 Zn(X(i)) + Zn(X

(i)A))2

1N

∑Ni=1 Zn(X(i))2 −

(1

2N

∑Ni=1 Zn(X(i)) + Zn(X

(i)A))2 , (57)

where (X(i), X(i)A)i=1,...,N is a N -sample from the random variable (X,X∼A).

In particular, this approach avoids to compute the integrals presented in Oakley and O’Hagan [42] and

thus simplify the estimation of VA and V . Furthermore, it takes into account their joint distribution.

Remark. This result can easily be extended to the total Sobol’ index Stoti =

∑A⊃ i

SA. The reader is

referred to Sobol et al. [52] and Variance-based sensitivity analysis: Theory and estimation algorithms in

this handbook for examples of pick-freeze estimates of StotA .

3.5 Numerical estimates of Sobol’ indices by Gaussian process sampling

The sensitivity index SA,N (57) is obtained after substituting the Gaussian process Zn(x) for the original

computational model G(x). Therefore, it is a random variable defined on the same probability space as

Zn(x). The aim of this section is to present a simple methodology to get a sample SA,N of SA. From

this sample, an estimate of SA (56) and a quantification of its uncertainty can be deduced.

Sampling from the Gaussian predictive distribution To obtain a realization of SA,N , one has to

obtain a sample of Zn(x) on (X(i), X(i)A)i=1,...,N and then use Eq. (57). To deal with large N , an efficient

strategy is to sample Zn(x) using the Kriging conditioning method, see for example Chiles and Delfiner

[18]. Consider first the unconditioned, zero-mean Gaussian process:

Z(x) = GP (0, k(x,x′)) . (58)

Then, the Gaussian process:

Zn(x) = mn(x)− mn(x) + Z(x), (59)

where mn(x) = fT(x)β + rT(x)R−1(Z(X )− Fβ

)and β =

(FTR−1F

)−1FTR−1Z(X ) has the same

distribution as Zn(x). Therefore, one can compute realizations of Zn(x) from realizations of Z(x). Since

18

Page 22: Metamodel-based sensitivity analysis: polynomial … · Metamodel-based sensitivity analysis: polynomial ... Metamodel-based sensitivity analysis: polynomial chaos expansions

Z(x) is not conditioned, the problem is numerically easier. Among the available Gaussian process sam-

pling methods, several can be mentioned: Cholesky decomposition [43], Fourier spectral decomposition

[54], Karhunen-Loeve spectral decomposition [43] and the propagative version of the Gibbs sampler [32].

Remark. Let suppose that a new point x(n+1) is added to the experimental design X . A classical re-

sult of conditional probability implies that the new predictive distribution [Z(x)|Z(X ) = Y, Z(x(n+1)) =

G(x(n+1)), σ2,θ] is identical to [Zn(x)|Zn(x(n+1)) = G(x(n+1)), σ2,θ]. Therefore, Zn(x) can be viewed as

an unconditioned Gaussian process and, using the Kriging conditioning method, realizations of [Z(x)|Z(X ) =

Y, Z(x(n+1)) = G(x(n+1)), σ2,θ] can be derived from realizations of Zn(x) using the following equation:

Zn+1(x) =kn(x(n+1),X )

kn(x(n+1),x(n+1))

(G(x(n+1))− Zn(x(n+1))

)+ Zn(x). (60)

Therefore, it is easy to calculate a new sample of SA,N after adding a new point x(n+1) to the experimental

design set X . This result is used in the R CRAN package “sensitivity” to perform sequential design for

sensitivity analysis using a Stepwise Uncertainty Reduction (SUR) strategy [5].

3.5.1 Meta-model and Monte-Carlo sampling errors

Let us denote bySNA,i, i = 1, . . . ,m

a sample set of SA,N (57) where of size m > 0. From this sample

set, the following unbiased estimate of SA can be deduced:

SA =1

m

m∑

i=1

SNA,i. (61)

with variance:

σ2SA

=1

m− 1

m∑

i=1

(SNA,i − SA

)2

. (62)

The term σ2SA

represents the uncertainty on the estimate of SA (56) due to the metamodel approximation.

Therefore, with the presented strategy, one can both obtain an unbiased estimate of the sensitivity index

SA and a quantification of its uncertainty.

Finally it may be of interest to evaluate the error due to the pick-freeze approximation and to compare it

to the error due to the metamodel. To do so, one can use the central limit theorem [31; 17] or a bootstrap

procedure [34]. In particular, a methodology to evaluate the uncertainty on the sensitivity index due to

both the Gaussian process and to the pick-freeze approximations is presented in Le Gratiet et al. [34]. It

makes it possible to determine the value of N such that the pick-freeze approximation error is negligible

compared to that of the metamodel.

3.6 Summary

Gaussian Process regression makes it possible to perform sensitivity analysis on complex computational

models using a limited number of model evaluations. An important feature of this method is that one

can propagate the Gaussian process approximation error to the sensitivity index estimates. This allows

the construction of sequential design strategies optimized for sensitivity analysis. It also provides a

19

Page 23: Metamodel-based sensitivity analysis: polynomial … · Metamodel-based sensitivity analysis: polynomial ... Metamodel-based sensitivity analysis: polynomial chaos expansions

powerful tool to visualize the main effect of a group of variables and the uncertainty of its estimate.

Another advantage of this approach is that Gaussian process regression has been thoroughly investigated

in the literature and can be used in various problems. For example, the method can be adapted for

non-stationary numerical models by using a treed Gaussian process as in Gramacy and Taddy [27].

Furthermore, it can also be used for multifidelity computer codes, i.e. codes which can be run at multiple

level of accuracy (see Le Gratiet et al. [34]).

4 Applications

In this section, metamodel-based sensitivity analysis is illustrated on several academic and engineering

examples.

4.1 Ishigami function

The Ishigami function is given by:

G(x1, x2, x3) = sin(x1) + 7 sin(x2)2 + 0.1x43 sin(x1). (63)

The input distributions of X1, X2 and X3 are uniform over the interval [−π, π]3. This is a classical

academic benchmark for sensitivity analysis, with first-order Sobol’ indices:

S1 = 0.3138 S2 = 0.4424 S3 = 0. (64)

To compare polynomial chaos expansions and Gaussian process modeling on this example, experimental

designs of different sizes n are considered. For each size n, 100 Latin Hypercube Sampling sets (LHS)

are computed so as to replicate the procedure and assess statistical uncertainty.

For the polynomial chaos approach, the coefficients are calculated based on a degree-adaptive LARS

strategy (for details, see Blatman and Sudret [13]), resulting in a sparse basis set. The maximum

polynomial degree is adaptively selected in the interval 3 ≤ p ≤ 15 based on LOO cross-validation

error estimates (see Eq. (27)).

For the Gaussian process approach, a tensorized Matern-5/2 covariance kernel is chosen (see Rasmussen

and Williams [43]) with trend functions given by:

fT(x) =

1 x2 x22 x3

1 x32 x4

1 x42

. (65)

The hyper-parameters θ are estimated with a Leave-One-Out cross validation procedure while the pa-

rameters β and σ2 are estimated with a restricted maximum likelihood method.

First we illustrate in Figure 6 the accuracy of the models with respect to the sample size n. The Nash-

Sutcliffe model efficiency coefficient (also called predictivity coefficient) is defined as follows:

Q2 = 1−∑ntest

i=1 (G(x(i))− G(x(i)))2

∑ntest

i=1 (G(x(i))− G)2, G =

1

ntest

ntest∑

i=1

G(x(i)), (66)

20

Page 24: Metamodel-based sensitivity analysis: polynomial … · Metamodel-based sensitivity analysis: polynomial ... Metamodel-based sensitivity analysis: polynomial chaos expansions

where G(x(i)) is the prediction given by the polynomial chaos or the Gaussian process regression model

on the ith point of a test sample of size ntest = 10, 000. This test sample set is randomly generated from

a uniform distribution. The closer Q2 is to 1, the more accurate the metamodel is.

40 60 80 100 120

0.2

0.4

0.6

0.8

1.0

Polynomial chaos

n

Q2

40 60 80 100 120

0.2

0.4

0.6

0.8

1.0

Gaussian process regression

n

Q2

Figure 6: Q2 coefficient as a function of the sample size n for the Ishigami function. For each n, the

box-plots represent the variations of Q2 obtained over 100 LHS replications.

We emphasize that checking the metamodel accuracy (see Figure 6) is very important since a metamodel-

based sensitivity analysis provides sensitivity indices for the metamodel and not for the true model G(x).

Therefore, the estimated indices are relevant only if the considered surrogate model is accurate.

Figure 7 shows the Sobol’ index estimates with respect to the sample size n. For the Gaussian process

regression approach, the convergence for is reached for n = 100. It corresponds to a Q2 coefficient greater

than 90%. Convergence of the PCE approach is somewhat faster, with comparable accuracy achieved

21

Page 25: Metamodel-based sensitivity analysis: polynomial … · Metamodel-based sensitivity analysis: polynomial ... Metamodel-based sensitivity analysis: polynomial chaos expansions

with n = 60 and almost perfect accuracy for n = 100. Therefore, the convergence of the estimators of

the Sobol’ indices in Eqs. (36) to (39) is expected to be comparable to that of Q2. Note that the PCE

approach also provides second order- and total Sobol’ indices for free, as shown in Sudret [59].

40 60 80 100 120

0.0

0.2

0.4

0.6

0.8

1.0

Polynomial chaos

n

Sob

ol

40 60 80 100 120

0.0

0.2

0.4

0.6

0.8

1.0

40 60 80 100 120

0.0

0.2

0.4

0.6

0.8

1.0

40 60 80 100 120

0.0

0.2

0.4

0.6

0.8

1.0

Gaussian process regression

n

Sob

ol

40 60 80 100 120

0.0

0.2

0.4

0.6

0.8

1.0

40 60 80 100 120

0.0

0.2

0.4

0.6

0.8

1.0

Figure 7: First-order Sobol’ index estimates as a function of the sample size n for the Ishigami function.

The horizontal solid lines represent the exact values of S1, S2 and S3. For each n, the box-plot represents

the variations obtained from 100 LHS replications. The validation set comprises ntest = 10, 000 samples.

22

Page 26: Metamodel-based sensitivity analysis: polynomial … · Metamodel-based sensitivity analysis: polynomial ... Metamodel-based sensitivity analysis: polynomial chaos expansions

4.2 G-Sobol function

The G-Sobol function is given by :

G(x) =d∏

i=1

|4xi − 2|+ ai1 + ai

, ai ≥ 0. (67)

To benchmark the described metamodel-based sensitivity analysis methods in higher dimension, we select

d = 15. The exact first-order Sobol’ indices Si are given by the following equations:

Vi =1

3(1 + ai)2, i = 1, . . . , d,

V =d∏

i=1

(1 + Vi)− 1,

Si = Vi/V.

(68)

In this example, vector a = a1, a2, . . . , ad is equal to:

a = 1, 2, 5, 10, 20, 50, 100, 500, 1000, 1000, 1000, 1000, 1000, 1000, 1000 . (69)

As in the previous section, different sample sizes n are considered and 100 LHS replications are computed

for each n. Sparse polynomial chaos expansions are obtained with the same strategy as for the Ishigami

function: adaptive polynomial degree selection with 3 < p < 15 and LARS-based calculation of the

coefficients. For the Gaussian process regression model, a tensorized Matern-5/2 covariance kernel is

considered with a constant trend function f(x) = 1. The hyper-parameter θ is estimated with a Leave-

One-Out cross validation procedure and the parameters β and σ2 are estimated with the maximum

likelihood method.

The accuracy of the metamodels with respect to n is presented in Figure 8. It is computed from a test

sample set of size ntest = 10, 000. The convergence of the estimates of the first four first-order Sobol’

indices is represented in Figure 9. Both metamodel-based estimations yield excellent results already

with n = 100 samples in the experimental design. This is expected due to the good accuracy of both

metamodels for all the n considered (see Figure 8).

Finally, Table 2 provides the Sobol’ index estimates median and root mean square error for n = 100 and

n = 500. As presented in Figure 9, the estimates of the largest Sobol’ indices are very accurate. Note

that the remaining first order indices are insignificant. One can observe that the RMS error over the 100

LHS replications is slightly smaller when using PCE for both n = 100 and n = 500 ED points. Note that

the second order- and total Sobol’ indices are also available for free when using PCE.

23

Page 27: Metamodel-based sensitivity analysis: polynomial … · Metamodel-based sensitivity analysis: polynomial ... Metamodel-based sensitivity analysis: polynomial chaos expansions

100 200 300 400 500

0.80

0.85

0.90

0.95

1.00

Polynomial chaos

n

Q2

100 200 300 400 500

0.80

0.85

0.90

0.95

1.00

Gaussian process regression

n

Q2

Figure 8: Q2 coefficient as a function of the sample size n for G-Sobol academic example. For each n,

the box-plot represents the variations of Q2 obtained from 100 LHS.

24

Page 28: Metamodel-based sensitivity analysis: polynomial … · Metamodel-based sensitivity analysis: polynomial ... Metamodel-based sensitivity analysis: polynomial chaos expansions

100 200 300 400 500

0.0

0.2

0.4

0.6

0.8

1.0

Polynomial chaos

n

Sob

ol

100 200 300 400 500

0.0

0.2

0.4

0.6

0.8

1.0

100 200 300 400 500

0.0

0.2

0.4

0.6

0.8

1.0

100 200 300 400 500

0.0

0.2

0.4

0.6

0.8

1.0

100 200 300 400 500

0.0

0.2

0.4

0.6

0.8

1.0

Gaussian process regression

n

Sob

ol

100 200 300 400 500

0.0

0.2

0.4

0.6

0.8

1.0

100 200 300 400 500

0.0

0.2

0.4

0.6

0.8

1.0

100 200 300 400 500

0.0

0.2

0.4

0.6

0.8

1.0

Figure 9: Sobol’ index estimates with respect to the sample size n for G-Sobol function. The horizontal

solid lines represent the true values of S1, S2, S3 and S4. For each n, the box-plot represents the variations

obtained from 100 LHS.

25

Page 29: Metamodel-based sensitivity analysis: polynomial … · Metamodel-based sensitivity analysis: polynomial ... Metamodel-based sensitivity analysis: polynomial chaos expansions

Table 2: Sobol’ index estimates for the G-Sobol function. The median and the root mean square error

(RMSE) of the estimates are given for n = 100 and n = 500.

Polynomial chaos expansion Gaussian process regression

Median RMSE Median RMSE

Index Value 100 500 100 500 100 500 100 500

S1 0.604 0.619 0.607 0.034 0.007 0.618 0.599 0.035 0.012

S2 0.268 0.270 0.269 0.027 0.005 0.233 0.245 0.046 0.026

S3 0.067 0.063 0.065 0.014 0.003 0.045 0.070 0.029 0.016

S4 0.020 0.014 0.019 0.008 0.001 0.008 0.023 0.018 0.013

S5 0.005 0.002 0.005 0.003 0.001 8.6×10−4 1.8×10−3 0.014 0.013

S6 0.001 0.000 7.2×10−4 0.001 3.5×10−4 6.4×10−4 5.3×10−4 0.013 0.013

S7 0.000 0.000 1.1×10−4 1.1×10−3 1.4×10−4 5.3×10−4 3.0×10−4 0.013 0.013

S8 0.000 0.000 0.000 3.3×10−4 1.7×10−5 6.5×10−4 7.1×10−4 0.013 0.013

S9 0.000 0.000 0.000 4.1×10−4 1.1×10−5 8.5×10−4 4.4×10−4 0.14 0.013

S10 0.000 0.000 0.000 2.4×10−4 1.1×10−5 2.2×10−4 1.7×10−4 0.013 0.013

S11 0.000 0.000 0.000 9.5 ×10−4 1.2×10−5 5.5×10−4 -9.9×10−5 0.013 0.013

S12 0.000 0.000 0.000 5.2×10−4 2.1 ×10−5 2.6×10−4 4.1×10−4 0.013 0.013

S13 0.000 0.000 0.000 5.1×10−4 5.9 ×10−6 9.8×10−4 4.7×10−4 0.013 0.013

S14 0.000 0.000 0.000 8.8×10−4 1.9 ×10−5 1.8×10−4 6.9×10−4 0.013 0.013

S15 0.000 0.000 0.000 8.6×10−4 9.7×10−6 7.2×10−4 3.1×10−4 0.013 0.013

26

Page 30: Metamodel-based sensitivity analysis: polynomial … · Metamodel-based sensitivity analysis: polynomial ... Metamodel-based sensitivity analysis: polynomial chaos expansions

4.3 Morris function

The Morris function is given by [49]:

G(x) =

20∑

i=1

βiwi +

20∑

i<j

βijwiwj +

20∑

i<j<l

βijlwiwjwl + 5w1w2w3w4 (70)

where Xi ∼ U(0, 1), i = 1, . . . , 20 and wi = 2(xi − 1/2) for all i except for i = 3, 5, 7 where wi =

2(1.1xi/(xi + 0.1)− 1/2). The coefficients are defined as βi = 20, i = 1, . . . , 10; βij = −15, i, j = 1, . . . , 6;

βijl = −10, i, j, l = 1, . . . , 5. The remaining coefficients are set equal to βi = (−1)i and βij = (−1)i+j

and all the rest are zero. The reference values of the first-order Sobol’ indices of the Morris function are

calculated by a large Monte Carlo-based sensitivity analysis (n = 106).

As in the previous section different sample sizes n are considered and 100 LHS replications are computed

for each n. Sparse polynomial chaos expansions are obtained by adaptive polynomial degree selection

5 < p < 13 and LARS-based calculation of the coefficients.

The accuracy of the metamodels with respect to n is presented in Figure 10. It is computed from a

test sample of size ntest = 10, 000. As expected due to the complexity and dimensionality of the Morris

function, both metamodels show a slower overall convergence rate with the number of samples with respect

to the previous examples. Polynomial chaos expansions show in this case remarkably more scattering in

their performance for smaller experimental designs with respect to Gaussian process regression. This is

likely due to the comparatively large amount of prior information in the form of trend functions provided

to the Gaussian process, not used for PCE.

The convergence of the estimates of three selected first-order Sobol’ indices (the largest S9 and two

intermediate ones S3 and S8) is represented in Figure 11. Both methods perform very well with as

few as 250 samples. PCE, however, shows a more standard convergence behavior both in mean value in

dispersion. Gaussian process regression retrieves the Sobol’ estimates very accurately even with extremely

small experimental designs, but no clear convergence pattern can be seen for larger datasets.

Finally, Table 3 provides the a detailed breakdown of the Sobol’ index estimates, including median and

root mean square error (RMSE), for n = 100 and n = 500.

27

Page 31: Metamodel-based sensitivity analysis: polynomial … · Metamodel-based sensitivity analysis: polynomial ... Metamodel-based sensitivity analysis: polynomial chaos expansions

100 200 300 400 500

0.2

0.4

0.6

0.8

1.0

Polynomial chaos

n

Q2

100 200 300 400 500

0.0

0.2

0.4

0.6

0.8

1.0

Gaussian process regression

n

Q2

Figure 10: Q2 coefficient as a function of the sample size n for the Morris function example. For each n,

the box-plot represents the variations of Q2 obtained from 100 LHS.

28

Page 32: Metamodel-based sensitivity analysis: polynomial … · Metamodel-based sensitivity analysis: polynomial ... Metamodel-based sensitivity analysis: polynomial chaos expansions

100 200 300 400 500

0.0

0.2

0.4

0.6

0.8

1.0

Polynomial chaos

n

Sob

ol

100 200 300 400 500

0.0

0.2

0.4

0.6

0.8

1.0

100 200 300 400 500

0.0

0.2

0.4

0.6

0.8

1.0

100 200 300 400 500

0.0

0.2

0.4

0.6

0.8

1.0

Gaussian process regression

n

Sob

ol

100 200 300 400 500

0.0

0.2

0.4

0.6

0.8

1.0

100 200 300 400 500

0.0

0.2

0.4

0.6

0.8

1.0

Figure 11: First-order Sobol’ index estimates as a function of the sample size n for the Morris function.

The horizontal solid lines represent the exact values of S3, S8 and S9. For each n, the box-plot represents

the variations obtained from 100 LHS replications.

29

Page 33: Metamodel-based sensitivity analysis: polynomial … · Metamodel-based sensitivity analysis: polynomial ... Metamodel-based sensitivity analysis: polynomial chaos expansions

Table 3: First-order Sobol’ indices estimation for the Morris function. The median and the root mean

square error (RMSE) of the estimates are given for n = 100 and n = 500.

Polynomial chaos expansion Gaussian process regression

Median RMSE Median RMSE

Index Value 100 500 100 500 100 500 100 500

S2 0.005 0.000 0.005 0.252 0.017 0.011 0.004 0.109 0.085

S3 0.008 0.000 0.009 0.175 0.027 0.006 0.007 0.089 0.088

S1 0.017 0.000 0.015 0.304 0.047 0.003 0.017 0.130 0.109

S4 0.009 0.000 0.009 0.119 0.023 0.017 0.011 0.130 0.097

S5 0.016 0.000 0.015 0.230 0.043 0.005 0.016 0.120 0.109

S6 0.000 0.000 0.000 0.061 0.003 0.000 0.000 0.061 0.070

S7 0.069 0.045 0.068 0.585 0.058 0.072 0.062 0.095 0.123

S8 0.100 0.128 0.107 0.950 0.105 0.105 0.116 0.108 0.211

S9 0.150 0.192 0.160 1.241 0.143 0.127 0.143 0.246 0.117

S10 0.100 0.133 0.106 0.875 0.092 0.138 0.111 0.404 0.155

S11 0.000 0.000 0.000 0.185 0.003 0.004 0.000 0.088 0.074

S12 0.000 0.000 0.000 0.083 0.004 0.000 0.000 0.064 0.077

S13 0.000 0.000 0.000 0.081 0.003 0.000 0.000 0.064 0.074

S14 0.000 0.000 0.000 0.020 0.003 0.000 0.000 0.070 0.078

S15 0.000 0.000 0.000 0.140 0.003 0.000 0.000 0.065 0.075

S16 0.000 0.000 0.000 0.040 0.005 0.001 0.000 0.077 0.074

S17 0.000 0.000 0.000 0.264 0.004 0.000 0.000 0.065 0.075

S18 0.000 0.000 0.000 0.084 0.004 0.000 0.000 0.064 0.075

S19 0.000 0.000 0.000 0.083 0.004 0.000 0.000 0.064 0.076

S20 0.000 0.000 0.000 0.049 0.004 0.000 0.000 0.064 0.075

30

Page 34: Metamodel-based sensitivity analysis: polynomial … · Metamodel-based sensitivity analysis: polynomial ... Metamodel-based sensitivity analysis: polynomial chaos expansions

4.4 Maximum deflection of a truss structure

Sensitivity analysis is also of great interest for engineering models whose input parameters may have

different distributions. As an example consider the elastic truss structure depicted in Figure 12 (see

e.g. Blatman and Sudret [10]). This truss is made of two types of bars, namely horizontal bars with

cross-section A1 and Young’s modulus (stiffness) E1 on the one hand oblique bars with cross-section A2

and Young’s modulus (stiffness) E2 on the other hand. The truss is loaded with six vertical loads applied

on the top chord. Of interest is the maximum vertical displacement (called deflection) at mid-span. This

quantity is computed using a finite element model comprising elastic bar elements.

6 x 4m

2m

P1 P2 P3 P4 P5 P6

u4

Figure 12: Model of a truss structure with 23 members. The quantity of interest is the maximum

displacement at mid-span u4.

The various parameters describing the behavior of this truss structure are modeled by independent

random variables that account for the uncertainty in both the physical properties of the structure and

the applied loads. Their distributions are gathered in Table 4.

Table 4: Probabilistic input model of the truss structure

Variable Distribution Mean Standard Deviation

E1, E2 (Pa) Lognormal 2.1× 1011 2.1× 1010

A1 (m2) Lognormal 2.0× 10−3 2.0× 10−4

A2 (m2) Lognormal 1.0× 10−3 1.0× 10−4

P1-P6 (N) Gumbel 5.0× 104 7.5× 103

These input variables are collected in the random vector

X = E1, E2, A1, A2, P1, . . . , P6 . (71)

Using this notation, the maximal deflection of interest is cast as:

u4 = GFE(X). (72)

Different sparse polynomial chaos expansions are calculated assuming a maximal degree 3 < p < 10 using

LARS and the best expansion (in terms of smallest LOO error) is retained. For the Gaussian process

regression model, a tensorized Matern-5/2 covariance kernel is considered with a constant trend function

f(x) = 1. The hyper-parameter θ is estimated with a Leave-One-Out cross validation procedure and the

parameters β and σ2 are estimated with the maximum likelihood method.

31

Page 35: Metamodel-based sensitivity analysis: polynomial … · Metamodel-based sensitivity analysis: polynomial ... Metamodel-based sensitivity analysis: polynomial chaos expansions

The first order Sobol’ indices obtained from PCE and GP metamodels are reported in Table 5 in the case

when the experimental design is of size 100. In decreasing importance order, the important variables are

the properties of the chords (horizontal bars), then the loads close to mid-span, namely P3 and P4. The

Sobol’ indices of the latter are identical due to the symmetry of the model. Then come the loads P2 and

P5. The other variables (the loads P1 and P6 and the properties of the oblique bars) appear unimportant.

Table 5: Truss structure – First order Sobol’ indices

Variable Reference PCE Gaussian Process

A1 0.365 0.366 0.384

E1 0.365 0.369 0.362

P3 0.075 0.078 0.075

P4 0.074 0.076 0.069

P5 0.035 0.036 0.029

P2 0.035 0.036 0.028

A2 0.011 0.012 0.015

E2 0.011 0.012 0.008

P6 0.003 0.005 0.002

P1 0.002 0.005 0.000

The estimates of the three largest first-order Sobol’ indices which correspond to variables E1, P3 and P5

obtained for various sizes n of the LHS experimental design are plotted in Figure 13 as a function of n.

The reference solution is obtained by Monte-Carlo sampling with a sample set of size 6,000,000. Both

PCE- and GP-based Sobol’ indices converge to stable estimates as soon as n ≥ 60.

5 Conclusions

Sobol’ indices are recognized as good descriptors of the sensitivity of the output of a computational

model to its various input parameters. Classical estimation methods based on Monte Carlo simulation

are computationally expensive though. The required costs, in the order of 103 − 104 model evaluations,

are often not compatible with the advanced simulation models encountered in engineering applications.

For this reason, surrogate models may be first built up from a limited number of runs of the computational

model (the so-called experimental design), and the sensitivity analysis is then carried out by substituting

the surrogate model for the original one.

Polynomial chaos expansions and Gaussian processes are two popular methods that can be used for this

purpose. The advantage of the PCE approach is that the Sobol’ indices at any order may be computed

analytically once the expansion is available. In this contribution, least-square minimization techniques

are presented to compute the PCE coefficients, yet any intrusive or non intrusive method could be used

as an alternative.

32

Page 36: Metamodel-based sensitivity analysis: polynomial … · Metamodel-based sensitivity analysis: polynomial ... Metamodel-based sensitivity analysis: polynomial chaos expansions

Polynomial chaos expansion

n

Sob

ol

20 40 60 80 100 120 140 1000

0.0

0.1

0.2

0.3

0.4

0.5

Gaussian process regression

n

Sob

ol

0.0

0.1

0.2

0.3

0.4

0.5

20 40 60 80 100 120 140 1000

Figure 13: Truss structure – First-order Sobol’ index estimates as a function of the sample size n for the

truss model. The horizontal solid lines represent reference values of input variables E1, P3 and P5 from

a Monte Carlo estimate on 6,000,000 samples.

In contrast Gaussian process surrogate models are used together with Monte Carlo simulation for estimat-

ing the Sobol’ indices. The advantage of this approach is that the metamodel error can be included in the

estimators. Note that bootstrap techniques can be used similarly to calculate and include metamodeling

error also for PCE-based sensitivity analysis, as demonstrated by Dubreuil et al. [22].

As shown in the various comparisons, PCE and GP give similar accuracy (measured in terms of the Q2

validation coefficient) for a given size of the experimental design in a broad range of applications. The

replication of the analyses with different random designs of the same size show a smaller scatter using GP

for extremely small designs, whereas PCE becomes more stable for medium-size designs. Selecting the

33

Page 37: Metamodel-based sensitivity analysis: polynomial … · Metamodel-based sensitivity analysis: polynomial ... Metamodel-based sensitivity analysis: polynomial chaos expansions

best technique is in the end problem-dependent, and it is worth comparing the two approaches using the

same experimental design, as it can be done in recent sensitivity analysis toolboxes such as OpenTURNS

[2] and UQLab [36].

Finally it is worth mentioning that the so-called derivative-based global sensitivity measures (DGSM)

originally introduced by Sobol’ and Kucherenko [51] can also be computed using surrogate models. In

particular, polynomial chaos expansions may be used to compute the DGSM analytically, as shown in

Sudret and Mai [61]. The recent combination of polynomial chaos expansions and Gaussian processes

into PC-Kriging [48] also appears promising for estimating sensitivity indices from extremely small ex-

perimental designs.

References

[1] Abramowitz, M. and I. Stegun (1970). Handbook of mathematical functions. Dover Publications, Inc.

[2] Andrianov, G., S. Burriel, S. Cambier, A. Dutfoy, I. Dutka-Malen, E. de Rocquigny, B. Sudret,

P. Benjamin, R. Lebrun, F. Mangeant, and M. Pendola (2007). Open TURNS, an open source initiative

to Treat Uncertainties, Risks’N Statistics in a structured industrial approach. In Proc. ESREL’2007

Safety and Reliability Conference, Stavenger, Norway.

[3] Bachoc, F. (2013). Cross validation and maximum likelihood estimations of hyper-parameters of

gaussian processes with model misspecification. Computational Statistics & Data Analysis 66, 55–69.

[4] Bates, R. A., R. Buck, E. Riccomagno, and H. Wynn (1996). Experimental design and observation

for large systems. Journal of the Royal Statistical Society, Series B 58 (1), 77–94.

[5] Bect, J., D. Ginsbourger, L. Li, V. Picheny, and E. Vazquez (2012). Sequential design of computer

experiments for the estimation of a probability of failure. Statistics and Computing 22, 773–793.

[6] Berveiller, M., B. Sudret, and M. Lemaire (2004). Presentation of two methods for computing the

response coefficients in stochastic finite element analysis. In Proc. 9th ASCE Specialty Conference on

Probabilistic Mechanics and Structural Reliability, Albuquerque, USA.

[7] Berveiller, M., B. Sudret, and M. Lemaire (2006). Stochastic finite elements: a non intrusive approach

by regression. Eur. J. Comput. Mech. 15 (1-3), 81–92.

[8] Bieri, M. and C. Schwab (2009). Sparse high order FEM for elliptic sPDEs. Comput. Methods Appl.

Mech. Engrg 198, 1149–1170.

[9] Blatman, G. (2009). Adaptive sparse polynomial chaos expansions for uncertainty propagation and

sensitivity analysis. Ph. D. thesis, Universite Blaise Pascal, Clermont-Ferrand.

[10] Blatman, G. and B. Sudret (2008). Sparse polynomial chaos expansions and adaptive stochastic

finite elements using a regression approach. Comptes Rendus Mecanique 336 (6), 518–523.

34

Page 38: Metamodel-based sensitivity analysis: polynomial … · Metamodel-based sensitivity analysis: polynomial ... Metamodel-based sensitivity analysis: polynomial chaos expansions

[11] Blatman, G. and B. Sudret (2010a). An adaptive algorithm to build up sparse polynomial chaos

expansions for stochastic finite element analysis. Prob. Eng. Mech. 25, 183–197.

[12] Blatman, G. and B. Sudret (2010b). Efficient computation of global sensitivity indices using sparse

polynomial chaos expansions. Reliab. Eng. Sys. Safety 95, 1216–1229.

[13] Blatman, G. and B. Sudret (2011). Adaptive sparse polynomial chaos expansion based on Least

Angle Regression. J. Comput. Phys. 230, 2345–2367.

[14] Brown, S., J. Beck, H. Mahgerefteh, and E. Fraga (2013). Global sensitivity analysis of the impact

of impurities on CO2 pipeline failure. Reliab. Eng. Sys. Safety 115, 43–54.

[15] Buzzard, G. (2012). Global sensitivity analysis using sparse grid interpolation and polynomial chaos.

Reliab. Eng. Sys. Safety 107, 82–89.

[16] Buzzard, G. and D. Xiu (2011). Variance-based global sensitivity analysis via sparse-grid interpola-

tion and cubature. Comm. Comput. Phys. 9 (3), 542–567.

[17] Chastaing, G. and L. Le Gratiet (2015). Anova decomposition of conditional gaussian processes for

sensitivity analysis with dependent inputs. Journal of Statistical Computation and Simulation 85 (11),

2164–2186.

[18] Chiles, J. and P. Delfiner (1999). Geostatistics: modeling spatial uncertainty. Wiley series in

probability and statistics (Applied probability and statistics section).

[19] Crestaux, T., O. Le Maıtre, and J.-M. Martinez (2009). Polynomial chaos expansion for sensitivity

analysis. Reliab. Eng. Sys. Safety 94 (7), 1161–1172.

[20] Ditlevsen, O. and H. Madsen (1996). Structural reliability methods. J. Wiley and Sons, Chichester.

[21] Doostan, A. and H. Owhadi (2011). A non-adapted sparse approximation of pdes with stochastic

inputs. J. Comput. Phys. 230 (8), 3015–3034.

[22] Dubreuil, S., M. Berveiller, F. Petitjean, and M. Salaun (2014). Construction of bootstrap confidence

intervals on sensitivity indices computed by polynomial chaos expansion. Reliab. Eng. Sys. Safety 121,

263–275.

[23] Dubrule, O. (1983). Cross validation of kriging in a unique neighborhood. Mathematical Geology 15,

687–699.

[24] Efron, B., T. Hastie, I. Johnstone, and R. Tibshirani (2004). Least angle regression. Annals of

Statistics 32, 407–499.

[25] Fajraoui, N., F. Ramasomanana, A. Younes, T. Mara, P. Ackerer, and A. Guadagnini (2011). Use of

global sensitivity analysis and polynomial chaos expansion for interpretation of nonreactive transport

experiments in laboratory-scale porous media. Water Resources Research 47 (2).

35

Page 39: Metamodel-based sensitivity analysis: polynomial … · Metamodel-based sensitivity analysis: polynomial ... Metamodel-based sensitivity analysis: polynomial chaos expansions

[26] Ghanem, R. and P. Spanos (1991). Stochastic finite elements – A spectral approach. Springer Verlag,

New York. (Reedited by Dover Publications, Mineola, 2003).

[27] Gramacy, R. and M. Taddy (2012). Categorical inputs, sensitivity analysis, optimization and im-

portance tempering with tgp version 2, an r package for treed gaussian process models. Journal of

Statistical Software 33, 1–48.

[28] Harville, D. (1977). Maximum likelihood approaches to variance component estimation and to related

problems. Journal of the American Statistical Association 72 (358), 320–338.

[29] Iooss, B. and P. Lemaıtre (2015). Uncertainty management in simulation-optimization of com-

plex systems: algorithms and applications, Chapter A review on global sensitivity analysis methods.

Springer.

[30] Jakeman, J., M. Eldred, and K. Sargsyan (2015). Enhancing `1-minimization estimates of polynomial

chaos expansions using basis selection. J. Comput. Phys. 289, 18–34.

[31] Janon, A., T. Klein, A. Lagnoux, M. Nodet, and C. Prieur (2014, 1). Asymptotic normality and

efficiency of two sobol index estimators. ESAIM: Probability and Statistics 18, 342–364.

[32] Lantuejoul, C. and N. Desassis (2012, June). Simulation of a Gaussian random vector: A propagative

version of the Gibbs sampler. In The 9th International Geostatistics Congress, Oslo., Oslo, Norway,

pp. 1747181.

[33] Le Gratiet, L. and C. Cannamela (2015). Cokriging-based sequential design strategies using fast

cross-validation techniques for multi-fidelity computer codes.

[34] Le Gratiet, L., C. Cannamela, and B. Iooss (2014). A bayesian approach for global sensitivity

analysis of (multifidelity) computer codes. SIAM/ASA Journal on Uncertainty Quantification 2 (1),

336–363.

[35] Lebrun, R. and A. Dutfoy (2009). An innovating analysis of the Nataf transformation from the

copula viewpoint. Prob. Eng. Mech. 24 (3), 312–320.

[36] Marelli, S. and B. Sudret (2014). UQLab: A framework for uncertainty quantification in Matlab.

In Vulnerability, Uncertainty, and Risk (Proc. 2nd Int. Conf. on Vulnerability, Risk Analysis and

Management (ICVRAM2014), Liverpool, United Kingdom), pp. 2554–2563.

[37] Marrel, A., B. Iooss, B. Laurent, and O. Roustant (2009). Calculations of Sobol indices for the

Gaussian process metamodel. Reliability Engineering and System Safety 94, 742–751.

[38] Marrel, A., B. Iooss, F. Van Dorpe, and E. Volkova (2008). An efficient methodology for modeling

complex computer codes with gaussian processes. Computational Statistics & Data Analysis 52 (10),

4731–4744.

36

Page 40: Metamodel-based sensitivity analysis: polynomial … · Metamodel-based sensitivity analysis: polynomial ... Metamodel-based sensitivity analysis: polynomial chaos expansions

[39] McKay, M. D., R. J. Beckman, and W. J. Conover (1979). A comparison of three methods for

selecting values of input variables in the analysis of output from a computer code. Technometrics 2,

239–245.

[40] Munoz Zuniga, M., S. Kucherenko, and N. Shah (2013). Metamodelling with independent and

dependent inputs. Comput. Phys. Comm. 184, 1570 –1580.

[41] Niederreiter, H. (1992). Random number generation and quasi-Monte Carlo methods. Society for

Industrial and Applied Mathematics, Philadelphia, PA, USA.

[42] Oakley, J. and A. O’Hagan (2004). Probabilistic sensitivity analysis of complex models a Bayesian

approach. Journal of the Royal Statitistical Society series B 66, part 3, 751–769.

[43] Rasmussen, C. and C. Williams (2006). Gaussian Processes for Machine Learning. Cambridge: MIT

Press.

[44] Robert, C. (2007). The Bayesian choice: from decision-theoretic foundations to computational im-

plementation. New York: Springer.

[45] Sandoval, E. H., F. Anstett-Collin, and M. Basset (2012). Sensitivity study of dynamic systems

using polynomial chaos. Reliab. Eng. Sys. Safety 104, 15–26.

[46] Santner, T., B. Williams, and W. Notz (2003). The Design and Analysis of Computer Experiments.

New York: Springer.

[47] Sargsyan, K., C. Safta, H. Najm, B. Debusschere, D. Ricciuto, and P. Thornton (2014). Dimensional-

ity reduction for complex models via Bayesian compressive sensing. Int. J. Uncertain. Quantificat. 4 (1),

63–93.

[48] Schobi, R., B. Sudret, and J. Wiart (2015). Polynomial-chaos-based Kriging. Int. J. Uncertainty

Quantification 5 (2), 171–193.

[49] Schoebi, R., B. Sudret, and J. Wiart (2015). Polynomial-Chaos-based Kriging.

[50] Sobol, I. (1993). Sensitivity estimates for non linear mathematical models. Mathematical Modelling

and Computational Experiments 1, 407–414.

[51] Sobol’, I. and S. Kucherenko (2009). Derivative based global sensitivity measures and their link with

global sensitivity indices. Math. Comput. Simul. 79 (10), 3009–3017.

[52] Sobol, I., S. Tarantola, D. Gatelli, S. Kucherenko, and W. Mauntz (2007). Estimating the approx-

imation error when fixing unessential factors in global sensitivity analysis. Reliability Engineering &

System Safety 92 (7), 957–960.

[53] Soize, C. and R. Ghanem (2004). Physical systems with random uncertainties: chaos representations

with arbitrary probability measure. SIAM J. Sci. Comput. 26 (2), 395–410.

[54] Stein, M. (1999). Interpolation of Spatial Data. New York: Springer Series in Statistics.

37

Page 41: Metamodel-based sensitivity analysis: polynomial … · Metamodel-based sensitivity analysis: polynomial ... Metamodel-based sensitivity analysis: polynomial chaos expansions

[55] Storlie, C., L. Swiler, J. Helton, and C. Sallaberry (2009). Implementation and evaluation of nonpara-

metric regression procedures for sensitivity analysis of computationally demanding models. Reliability

Engineering & System Safety 94 (11), 1735–1763.

[56] Sudret, B. (2006). Global sensitivity analysis using polynomial chaos expansions. In P. Spanos and

G. Deodatis (Eds.), Proc. 5th Int. Conf. on Comp. Stoch. Mech (CSM5), Rhodos, Greece.

[57] Sudret, B. (2007). Uncertainty propagation and sensitivity analysis in mechanical models – contri-

butions to structural reliability and stochastic spectral methods. Technical report, Universite Blaise

Pascal, Clermont-Ferrand, France. Habilitation a diriger des recherches (229 pages).

[58] Sudret, B. (2008). Global sensitivity analysis using polynomial chaos expansions. Reliab. Eng. Sys.

Safety 93, 964–979.

[59] Sudret, B. (2015). Polynomial chaos expansions and stochastic finite element methods, Chapter 6.

Risk and Reliability in Geotechnical Engineering. Taylor and Francis.

[60] Sudret, B. and Y. Caniou (2013). Analysis of covariance (ancova) using polynomial chaos expansions.

In G. Deodatis (Ed.), Proc. 11th Int. Conf. Struct. Safety and Reliability (ICOSSAR’2013), New York,

USA.

[61] Sudret, B. and C.-V. Mai (2015). Computing derivative-based global sensitivity measures using

polynomial chaos expansions. Reliab. Eng. Sys. Safety 134, 241–250.

[62] van Beers, W. and J. Kleijnen (2008). Customized sequential designs for random simulation ex-

periments: Kriging metamodelling and bootstrapping. European journal of operational research 186,

1099–1113.

[63] Welch, W. J., R. J. Buck, J. Sacks, H. P. Wynn, T. J. Mitchell, and M. D. Morris (1992). Screening,

predicting, and computer experiments. Technometrics 34 (1), 15–25.

[64] Xiu, D. and G. Karniadakis (2002). The Wiener-Askey polynomial chaos for stochastic differential

equations. SIAM J. Sci. Comput. 24 (2), 619–644.

[65] Younes, A., T. Mara, N. Fajraoui, F. Lehmann, B. Belfort, and H. Beydoun (2013). Use of global

sensitivity analysis to help assess unsaturated soil hydraulic parameters. Vadose Zone Journal 12 (1).

38